首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
To verify the validity of thermodynamic approaches to the prediction of cellular behavior, cell spreading of three different cell types on solid substrata was determined in vitro. Solid substrata as well as cell types were selected on the basis of their surface free energies, calculated from contact angle measurements. The surface free energies of the solid substrata ranged from 18–116 erg cm−2. To measure contact angles on cells, a technique was developed in which a multilayer of cells was deposited on a filter and air dried. Cell surface free energies ranged from 60 erg cm−2 for fibroblasts, and 57 for smooth muscle cells, to 91 for HeLa epithelial cells. After adsorption of serum proteins, cell surface free energies of all three cell types converged to approx 74 erg cm−2. The spreading of these cell types from RPMI 1640 medium on the various solid substrata showed that both in the presence and in the absence of serum proteins in the medium, cells spread poorly on low energy substrata (Y s <50 erg cm−2), whereas good cell spreading was observed on the higher energy substrata. Calculations of the interfacial free energy of adhesion (ΔF adh) show that ΔF adh decreases with increasingY s , and equals zero around 45 erg cm−2 for all three cell types in the presence of serum proteins and for HeLa epithelium cells in the absence of serum proteins. This explains the spreading of these cells on the various substrata upon a thermodynamic basis. The results clearly show that substratum surface free energy has a predictive value with respect to cell spreading in vitro, both in the presence and absence of serum proteins. It is noted, however, that interfacial thermodynamics fail to explain the behavior of fibroblasts and smooth muscle cells in the absence of serum proteins, most likely because of the relatively high surface charges of these two cell types.  相似文献   

2.
The rabbit Na+/glucose cotransporter (SGLT1) exhibits a presteady-state current after step changes in membrane voltage in the absence of sugar. These currents reflect voltage-dependent processes involved in cotransport, and provide insight on the partial reactions of the transport cycle. SGLT1 presteady-state currents were studied as a function of external Na+, membrane voltage V m , phlorizin and temperature. Step changes in membrane voltage—from the holding V h to test values, elicited transient currents that rose rapidly to a peak (at 3–4 msec), before decaying to the steady state, with time constants τ≈4–20 msec, and were blocked by phlorizin (K i ≈30 μm). The total charge Q was equal for the application of the voltage pulse and the subsequent removal, and was a function of V m . The Q-V curves obeyed the Boltzmann relation: the maximal charge Q max was 4–120 nC; V 0.5, the voltage for 50% Q max was −5 to +30 mV; and z, the apparent valence of the moveable charge, was 1. Q max and z were independent of V h (between 0 and −100 mV) and temperature (20–30°C), while increasing temperature shifted V 0.5 towards more negative values. Decreasing [Na+] o decreased Q max, and shifted V 0.5 to more negative voltages 9by −100 mV per 10-fold decrease in [Na+] o ). The time constant τ was voltage dependent: the τ-V relations were bell-shaped, with maximal τmax 8–20 msec. Decreasing [Na+] o decreased τmax, and shifted the τ-V curves towards more negative voltages. Increasing temperature also shifted the τ-V curves, but did not affect τmax. The maximum temperature coefficient Q 10 for τ was 3–4, and corresponds to an activation energy of 25 kcal/mole. Simulations of a 6-state ordered kinetic model for rabbit Na+/glucose cotransport indicate that charge-movements are due to Na+-binding/dissociation and a conformational change of the empty transporter. The model predicts that (i) transient currents rise to a peak before decay to steady-state; (ii) the τ-V relations are bell-shaped, and shift towards more negative voltages as [Na+] o is reduced; (iii) τmax is decreased with decreasing [Na+] o ; and (iv) the Q-V relations are shifted towards negative voltages as [Na+] o is reduced. In general, the kinetic properties of the presteady-state currents are qualitatively predicted by the model. Received: 12 August 1996/Revised: 30 September 1996  相似文献   

3.
The growth and production of the Baltic clam Macoma balthica in the southeastern part of the Baltic Sea were studied. The shell length of M. balthica reached 23.5 mm, the maximum age was 13+ years. The linear growth was described by the von Bertalanffy equation for shallow-water area (depths 9–40 m): L τ = 23.99(1 − e −0.1293(τ − (−0.9578))), and for the deep-water area (41–81 m): L τ = 20.61(1 − e −0.1813(τ − (−0.5608))). The annual production was lower (25.35 ± 1.72 kJ/m2) in the shallow-water area than in the deep-water area (71.23 ± 4.48 kJ/m2), with values of P s /B ratio 0.44 and 0.38, respectively.  相似文献   

4.
The patch clamp K+-conductance G of the nicotinic acetylcholine receptor (AcChoR) dimer (Mr≈ 590 000) of Torpedo californica, reconstituted in lipid vesicles, which decreases with increasing Ca2+-concentration in the range 0.1≤[Ca2+]/mM≤2, can be quantitatively rationalized by Ca2+-binding to negatively charged sites, causing charge reversal reducing the normal K+-accumulation in the channel vestibules. Cleavage of the sialic acid residues (up to 20±2 per dimer) reduces the K+-accumulation factor α = G0/G from α = 3±0.8 of the normal AcChoR to α = 2±0.7 for the desialyated AcChoR. Desialysation also decreases the Ca2+-sensitivity of the conductance from G0 = 96.6±6 pS at [Ca2+]→0 of the normal AcChoR to G0 = 84.2±6 pS. Endogenous hyperphosphorylation (to up to 28±4 phosphates per dimer) enhances the vestibular K+-accumulation to α = 3.6±0.7, without affecting the Ca2+-dissociation equilibrium constant KCa = 0.34± 0.05 mM at 295 K (22 °C). Most interestingly, even in the absence of AcCho, the hyperphosphorylated AcChoR dimer exhibits spontaneously long-lasting open channel events (τ = 200±50 ms). At [AcCho] = 2 μM there are two open states (τ 1 = 20±10 ms, τ 2 = 140±60 ms) whereas the normal AcChoR dimer has only one open state (τ = 6±4 ms). – Physiologically important is that (i) the sialic acid and phosphate residues render the AcChoR conductance sensitive to control by divalent ions and (ii) the channel behavior of the hyperphosphorylated AcChoR without AcCho appears to indicate pathophysiologically high phosphorylation activity of the cell leading, among others, to myasthenic syndromes. Received: 10 November 1997 / Revised version: 12 January 1998 / Accepted: 7 March 1998  相似文献   

5.
A stretch-activated (SA) Cl channel in the plasma membrane of the human mast cell line HMC-1 was identified in outside-out patch-clamp experiments. SA currents, induced by pressure applied to the pipette, exhibited voltage dependence with strong outward rectification (55.1 pS at +100 mV and an about tenfold lower conductance at −100 mV). The probability of the SA channel being open (P o) also showed steep outward rectification and pressure dependence. The open-time distribution was fitted with three components with time constants of τ1o = 755.1 ms, τ2o = 166.4 ms, and τ3o = 16.5 ms at +60 mV. The closed-time distribution also required three components with time constants of τ1c = 661.6 ms, τ2c = 253.2 ms, and τ3c = 5.6 ms at +60 mV. Lowering extracellular Cl concentration reduced the conductance, shifted the reversal potential toward chloride reversal potential, and decreased the P o at positive potentials. The SA Cl currents were reversibly blocked by the chloride channel blocker 4,4′-diisothiocyanatostilbene-2,2′-disulfonic acid (DIDS) but not by (Z)-1-(p-dimethylaminoethoxyphenyl)-1,2-diphenyl-1-butene (tamoxifen). Furthermore, in HMC-1 cells swelling due to osmotic stress, DIDS could inhibit the increase in intracellular [Ca2+] and degranulation. We conclude that in the HMC-1 cell line, the SA outward currents are mediated by Cl influx. The SA Cl channel might contribute to mast cell degranulation caused by mechanical stimuli or accelerate membrane fusion during the degranulation process.  相似文献   

6.
 The ligand DOTASA was designed and synthesized in the aim of obtaining a kinetically and thermodynamically stable Gd(III) chelate which, through its uncoordinated carboxylate function, will provide an efficient pathway to couple the complex to bio- or macromolecules without affecting the coordination pattern of DOTA. Furthermore, it allows us to study the influence of an extra carboxylate arm on the parameters determining proton relaxivity in comparison to the commercial agent [Gd(DOTA)(H2O)]. A combined variable-temperature 17O NMR, EPR and nuclear magnetic relaxation dispersion study on the Gd(III) chelate resulted in k 298 ex=(6.3±0.2)×106 s–1 for the water exchange rate and τ298 R=125±2 ps for the rotational correlation time. The slight increase in both k 298 ex and τ298 R, as compared to those for [Gd(DOTA)(H2O)], is attributed to the presence of the extra negative charge. The longer rotational correlation time results in a proton relaxivity of 5.03 mM–1 s–1 for [Gd(DOTASA)(H2O)]2–, which is approximately 30% higher than that for [Gd(DOTA)(H2O)]. The increased water exchange rate of [Gd(DOTASA)(H2O)]2– has no consequence for proton relaxivity since this latter is exclusively limited by fast rotation for both complexes. However, for slowly rotating macromolecular agents, which contain a covalently coupled DOTASA unit instead of a coupled DOTA, this increased exchange rate will have a significant positive effect. Received: 31 December 1998 / Accepted: 4 March 1999  相似文献   

7.
Depolarization-activated H+-selective currents were studied using whole-cell and excised-patch voltage clamp methods in human monocytic leukemia THP-1 cells, before and after being induced by phorbol ester to differentiate into macrophage-like cells. The H+ conductance, g H, activated slowly during depolarizing pulses, with a sigmoidal time course. Fitted by a single exponential following a delay, the activation time constant, τact was roughly 10 sec at threshold potentials, decreasing at more positive potentials. Tail currents upon repolarization decayed mono-exponentially at all potentials. The tail current time constant, τtail, was voltage dependent, decreasing with hyperpolarization from 2–3 sec at 0 mV to ∼200 msec at −100 mV. Surprisingly, although τact depended strongly on pH o , τtail was completely independent of pH o . H+ currents were inhibited by Zn2+. Increasing pH o or decreasing pH i shifted the voltage-activation relationship to more negative potentials, tending to activate the g H at any given voltage. Studied in excised, inside-out membrane patches, H+ currents were larger and activated much more rapidly at lower bath pH (i.e., pH i ). In THP-1 cells differentiated into macrophages, the H+ current density was reduced by one-half, and τact was slower by about twofold. The properties of H+ channels in THP-1 cells and in other macrophage-related cells are compared. Received: 19 September 1995/Revised: 14 March 1996  相似文献   

8.
Seasonal changes in abundance and distribution pattern of soil micro-arthropods were studied in connection with a few environmental factors in a Japanese cedar (Cryptomeria japonica D. Don) plantation. The soil arthropods were sampled from three different depths at intervals of two months for two years. Of the collected animals (total 51000–155000 m−2), Collembola (20000–76000 m−2), oribatid mites (19000–55000 m−2) and carnivorous mites (6200–21000 m−2) were the numerically dominant animal groups. Low seasonal variations in abundance indicated their seasonal stability in population levels. The trends in seasonal fluctuation were similar among these groups and between the two years, showing bimodal pattern with little peaks in early summer and winter. The pattern of seasonal fluctuation in abundance of carnivorous mites (P d) was significantly synchronized with that in the total abundance of Collembola and oribatid mites (P τ). Thus, the number-ratios (P d/P τ) were fairly constant, ranging from 0.10 to 0.25. Seasonal changes in vertical distribution of the three animal groups showed a similar pattern for both years. The downward migrations were shown to be more affected by low temperatures in winter accompanied by snow coverage rather than by the desiccation of the surface soil in summer. All the three groups demonstrated as a whole slightly aggregated patterns of horizontal distribution throughout the two years. Temporal increases in the patchiness indices were observed from summer to autumn when the moisture content of the surface soil was low.  相似文献   

9.
Protoplasts were isolated from embryogenic suspension cultures derived from avocado (Persea americana Mill.) zygotic embryos and nucellus in an enzyme digestion solution consisting of 1% cellulase Onozuka RS, 1% Macerase R10, 0.2% Pectolyase Y-23, 0.7 M mannitol. 24.5 mM CaCl2, 0.92 mM NaH2PO4 and 6.25 2-[N-morpholino]ethanesulfonic acid (1.5 ml) mixed with 0.7 M MS8P (2.5 ml). MS-8P medium consisted of Murashige and Skoog salts without NH4NO3, 1 mg l–1 thiamine HCl, 100 mg l–1 myo-inositol, 3.1 g l–1 glutamine and 8P organic addenda. Medium osmolarity was adjusted with 0.15 M sucrose and 0–0.55 M mannitol. Protoplast yields of 3.5×106 protoplasts g–1 were obtained. Growth and development of the protoplasts were significantly affected by osmolarity, nitrogen source, plating density and culture medium dilution. Under optimum conditions, proembryos developed directly from embryogenic protoplasts and subsequently into somatic embryos. Optimum conditions for somatic embryo development included the culture of protoplasts at a density of 0.8–1.6×105 ml–1 in 0.4 M MS8P for 2–3 weeks, followed by subculture in 0.15 M MS8P at a diluted density of 20–40× for 1 month in darkness to obtain somatic embryos. Mature somatic embryos were recovered on semisolid medium; however, a low frequency of plantlet recovery (≤1%) from protoplast-derived somatic embryos was observed. Received: 9 February 1998 / Revision received: 4 May 1998 / Accepted: 15 May 1998  相似文献   

10.
The efficiency of cell-free protein synthesis combined with combinatorial selective 15N-labelling provides a method for the rapid assignment of 15N-HSQC cross-peaks to the 19 different non-proline amino-acid types from five 15N-HSQC spectra. This strategy was explored with two different constructs of the C-terminal domain V of the τ subunit of the Escherichia coli DNA polymerase III holoenzyme, τC16 and τC14. Since each of the five 15N-HSQC spectra contained only about one third of the cross-peaks present in uniformly labelled samples, spectral overlap was much reduced. All 15N-HSQC cross-peaks of the backbone amides could be assigned to the correct amino-acid type. Availability of the residue-type information greatly assisted the evaluation of the changes in chemical shifts observed for corresponding residues in τC16 vs. those in τC14, and the analysis of the structure and mobility of the C-terminal residues present in τC16 but not in τC14.  相似文献   

11.
We have used two different approaches to determine hydrodynamic parameters for mucins secreted by guinea-pig tracheal epithelial cells in primary culture. Cells were cultured under conditions that promote mucous cell differentiation. Secreted mucins were isolated as the excluded fraction from a Sepharose CL-4B gel filtration column run under strongly dissociating conditions. Biochemical analysis confirmed the identity of the high molecular weight material as mucins. Analytical ultracentrifugation was used to study the physical properties of the purified mucins. The weight average molecular mass (M w ) for three different preparations ranged from 3.3×106 to 4.7×106 g/mol (corresponding to an average structure of 1 – 2 subunits), and the sedimentation coefficient from 25.5 to 35 S. Diffusion coefficients ranging from 4.5×10–8 to 6.4×10–8 cm2/s were calculated using the Svedberg equation. A polydispersity index (M z /M w ) of ∼1.4 was obtained. Diffusivity values were also determined by image analysis of mucin granule exocytosis captured by videomicroscopy. The time course of hydration and dissolution of mucin was measured and a relationship is presented which models both phases, each with first order kinetics, in terms of a maximum radius and rate constants for hydration and dissolution. A median diffusivity value of 8.05×10–8 cm2/s (inter-quartile range = 1.11×10–7 to 6.08×10–8 cm2/sec) was determined for the hydration phase. For the dissolution phase, a median diffusivity value of 6.98×10–9 cm2/s (inter-quartile range = 1.47×10–8 to 3.25×10–9 cm2/sec) was determined. These values were compared with the macromolecular diffusion coefficients (D 20,w ) obtained by analytical ultracentrifugation. When differences in temperature and viscosity were taken into account, the resulting D 37,g was within the range of diffusivity values for dissolution. Our findings show that the physicochemical properties of mucins secreted by cultured guinea-pig tracheal epithelial cells are similar to those of mucins of the single or double subunit type purified from respiratory mucus or sputum. These data also suggest that measurement of the diffusivity of dissolution may be a useful means to estimate the diffusion coefficient of mucins in mucus gel at the time of exocytosis from a secretory cell. Received: 10 March 1998 / Accepted: 27 March 1998  相似文献   

12.
The biophysical properties of the interaction between fibronectin and its membrane receptor were inferred from adhesion tests on living cells. Individual fibroblasts were maintained on fibronectin-coated glass for short time periods (1–16 s) using optical tweezers. After contact, the trap was removed quickly, leading to either adhesion or detachment of the fibroblast. Through a stochastic analysis of bond kinetics, we derived equations of adhesion probability versus time, which fit the experimental data well and were used to compute association and dissociation rates (k +=0.3–1.4 s−1 and k off=0.05–0.25 s−1, respectively). The bond distribution is binomial, with an average bond number ≤10 at these time scales. Increasing the fibronectin density (100–3000 molecules/μm2) raised k + in a diffusion-dependent manner, leaving k off relatively unchanged. Increasing the temperature (23–37 °C) raised both k + and k off, allowing calculation of the activation energy of the chemical reaction (around 20 k B T). Increasing the compressive force on the cell during contact (up to 60 pN) raised k + in a logarithmic manner, probably through an increase in the contact area, whereas k off was unaffected. Finally, by varying the pulling force to detach the cell, we could distinguish between two adhesive regimes, one corresponding to one bond, the other to at least two bonds. This transition occurred at a force around 20 pN, interpreted as the strength of a single bond. Received: 2 November 1999 / Revised version: 6 March 2000 / Accepted: 19 April 2000  相似文献   

13.
Summary In the neonatal rat lung, alveolar development occurs from postnatal Days 4–13, during which time there is a fourfold increase in interstitial fibroblasts. Factors influencing emergence of new septa and cell proliferation associated with septal elongation have yet to be identified, in part because of difficulties inherent in studying this process in vivo. Using flow cytometric analysis of the DNA content of freshly isolated lung fibroblasts, we found that proliferation, as indicated by the percentage of cells in S plus G2/M phases, peaked on postnatal Day 4 (P<0.04). By Days 9–10 the proliferation rate was lower than on Days 3, 4, 5, or 6 (P<0.005). We then evaluated rates of in vitro proliferation as a function of postnatal age in first passage fibroblasts and found that the proliferative phenotype expressed in vivo persists in vitro. Fibroblasts from 4–5-d-old pups increased in number and incorporated 3H-thymidine at a faster rate than did fibroblasts obtained from pups at other postnatal ages (P<0.0001). Age-dependent differences in cell cycle transit time were compared in fibroblasts synchronized by serum starvation and analyzed by flow cytometry at 2-h intervals from 13–21 h after release from serum starvation. A greater percentage of cells from 5-d-old pups entered S phase during this period than was seen for cells obtained from 2-, 9-, 13-, or 23-d-old rat pups (P=0.0001). Cells from 5-, 9-, and 13-d-old pups reentered G0/G1 by 21 h after release from serum starvation, in contrast to fibroblasts from 2- and 23-d-old rats which did not. Throughout the 15-h period after release from serum starvation, levels of cyclin E, which peaks at the G1/S border, were highest in the 5-d-old cells (P<0.025). Synchronization with 2.5 mM hydroxyurea which inhibits DNA synthesis completely abolished age-related differences in cell cycle transit time, implying that age-dependent differences in lung fibroblast proliferation rates are the result of events occurring before S-phase entry.  相似文献   

14.
The uptake of soluble phosphate by the green sulfur bacterium Chlorobium limicola UdG6040 was studied in batch culture and in continuous cultures operating at dilution rates of 0.042 or 0.064 h–1. At higher dilution rates, washout occurred at phosphate concentrations below 7.1 μM. This concentration was reduced to 5.1 μM when lower dilution rates were used. The saturation constant for growth on phosphate (K μ) was between 2.8 and 3.7 μM. The specific rates of phosphate uptake in continuous culture were fitted to a hyperbolic saturation model and yielded a maximum rate (Va max) of 66 nmol P (mg protein)–1 h–1 and a saturation constant for transport (K t) of 1.6 μM. In batch cultures specific rates of phosphate uptake up to 144 nmol P (mg protein)–1 h–1 were measured. This indicates a difference between the potential transport of cells and the utilization of soluble phosphate for growth, which results in a significant change in the specific phosphorus content. The phosphorus accumulated within the cells ranged from 0.4 to 1.1 μmol P (mg protein)–1 depending on the growth conditions and the availability of external phosphate. Transport rates of phosphate increased in response to sudden increases in soluble phosphate, even in exponentially growing cultures. This is interpreted as an advantage that enables Chl. limicola to thrive in changing environments. Received: 9 February 1998 / Accepted: June 1998  相似文献   

15.
The dynamics of the nucleobase and the ribose moieties in a 14-nt RNA cUUCGg hairpin-loop uniformly labeled with 13C and 15N were studied by 13C spin relaxation experiments. R1, R and the 13C-{1H} steady-state NOE of C6 and C1′ in pyrimidine and C8 and C1′ in purine residues were obtained at 298 K. The relaxation data were analyzed by the model-free formalism to yield dynamic information on timescales of pico-, nano- and milli-seconds. An axially symmetric diffusion tensor with an overall rotational correlation time τc of 2.31±0.13 ns and an axial ratio of 1.35±0.02 were determined. Both findings are in agreement with hydrodynamic calculations. For the nucleobase carbons, the validity of different reported 13C chemical shift anisotropy values (Stueber, D. and Grant, D. M., 2002 J. Am. Chem. Soc. 124, 10539–10551; Fiala et al., 2000 J. Biomol. NMR 16, 291–302; Sitkoff, D. and Case, D. A., 1998 Prog. NMR Spectroscopy 32, 165–190) is discussed. The resulting dynamics are in agreement with the structural features of the cUUCGg motif in that all residues are mostly rigid (0.82 < S2 < 0.96) in both the nucleobase and the ribose moiety except for the nucleobase of U7, which is protruding into solution (S2 = 0.76). In general, ribose mobility follows nucleobase dynamics, but is less pronounced. Nucleobase dynamics resulting from the analysis of 13C relaxation rates were found to be in agreement with 15N relaxation data derived dynamic information (Akke et al., 1997 RNA 3, 702–709). Electronic supplementary material Electronic supplementary material is available for this article at and accessible for authorised users.  相似文献   

16.
The seedling of an achlorophyllous orchid,Galeola septentrionalis, requires for its early growth anomalous atmospheric conditions appropriate to each process of development. The most favorable atmosphere for the enlargement of protocorm consists of 10% O2, 6% CO2 and 84% N2 at a pressure of 1.4 kg/cm2. Tolerable ranges of atmospheric pressure, concentrations of O2 and CO2 were 1.1–2.0 kg/cm2, 5–12% and 2–10%, respectively. Such a range for culture temperatures was 20–26 C. Subsequent development of the seedlings was little influenced by atmospheric pressure (1.0–1.8 kg/cm2) and concentration of CO2 (0–8%), but influenced by O2 concentration. Optimum and tolerable concentrations of O2 were 10–15% and 5–19.7%, respectively. These atmospheres are discussed hypothetically as the conditions required to pass an inevitable process specific to the epigenetic ontogeny of scobiform (sawdusty) seeds.  相似文献   

17.
Laboratory flume experiments were carried out, to measure the effect of biota on erodibility of mudflat sediments. The experiments sought to reproduce the environment of the lower mudflat at Hythe, Southampton Water, Southern England; this is characterised by fine grain-size and a surface layer of very fluid mud. Natural sediments were used to produce settled beds in the Lab Carousel, an annular flume of 2 m diameter. The following bed conditions were investigated diatom biofilms; the addition of cockles (Cerastoderma edule); and abiotic sediment, obtained by the addition of sodium hypochlorite. The erosion threshold (τcrit, calculated with the TKE method) was in the range 0.02–0.20 Pa. Bioconsolidation increased τcrit considerably: compared to the abiotic sediment experiment, τcrit was 5–10 times higher depending on the biofilm development. The relationship between τcrit and water content of sediment (the best proxy for sediment compaction) was as good, or better than between τcrit and chlorophyll a (proxy for biofilm development). When cockles were introduced, τcrit was significantly lower (reduction by 50–75% compared with the diatom biofilm experiments), reflecting the surface disturbance by the bivalves. The biofilm erosion was characterised by a patchy pattern: the bed surface stayed mainly uneroded and erosion was visible only on a few elongated patches commencing at some weakness points of the biofilm, then progressing downstream. The results illustrate the importance of the surface heterogeneity: the irregularities of a natural bed (weak points of the biofilm, bioturbations, microrelief, larger roughness elements like shells or algae, etc.) have a determinant effect on the erodibility of biofilms. Such characteristics may have more influence than biofilm strength, because the erosion starts from the weaker areas.  相似文献   

18.
 Dithionite has been found to reduce directly (without mediators) the Escherichia coli R2 subunit of ribonucleotide reductase. With dithionite (∼10 mM) in large excess, the reaction at 25  °C is complete in ∼10 h. Preparations of E. coli R2 have an FeIII 2 (met-R2) component in this work at ∼40% levels, alongside the fully active enzyme FeIII 2 . . . Tyr*, which has a tyrosyl radical at Tyr-122. In the pH range studied (7–8) the kinetics are biphasic. Rate laws for both phases give [S2O4 2–] and not [S2O4 2–]1/2 dependencies, and saturation kinetics are observed for the first time in R2 studies. No dependence on pH was detected. The kinetics (25  °C) of the first phase are reproduced in separate experiments using only met-R2, with association of S2O4 2– to met-R2, K=330 M–1, occurring prior to electron transfer, k et=4.8×10–4 s–1, I=0.100 M (NaCl). The second phase assigned to the reaction of FeIII 2 . . . Tyr* with S2O4 2– gives K=800 M–1 and k et=5.6×10–5 s–1. Bearing in mind the substantially smaller reduction potential for FeIII 2 compared to Tyr*, this is a quite remarkable finding, with implications similar to those already reported for the reaction of R2 with hydrazine, but with additional information provided by the saturation kinetics. The similarity in rates for the two phases (∼fourfold difference) suggests that reduction of FeIII 2 is occurring in both cases, and since S2O4 2– is involved a two-equivalent change is proposed with the formation of FeII 2 . . . Tyr* in the case of active R2. As a sequel to the second phase, intramolecular reduction of the strongly oxidising Tyr* by the FeII 2 is rapid, and further decay of FeIIFeIII is also fast. There is no stable mouse met-R2 form, and the single-phase reaction with dithionite gives saturation kinetics with K=208 M–1 and k et=1.7±10–3 s–1. Mechanistic implications, including the applicability of a pathway for electron transfer via FeA, are considered. Received: 25 February 1998 / Received: 20 August 1998  相似文献   

19.
The canopy water relations of old-growth Douglas-fir trees   总被引:4,自引:0,他引:4  
 We investigated whole tree water relations in 56–65 m tall, old-growth Pseudotsuga menziesii trees within the Wind River Canopy Crane site, Carson, Washington, USA. We measured at predawn and solar noon the vertical gradients in xylem pressure potential using a pressure chamber. On an Abies amabilis sapling located in the understory at the base of one of the study trees, predawn and solar noon xylem pressure potentials were also measured. Xylem pressure potential data were measured from late June through early September 1996 on foliage sampled from 1 to 64.5 m. Over this height gradient, predawn water potentials ranged from –0.23 to –1.10 MPa. Solar noon values showed an even greater range (from –0.44 to –2.51 MPa). At predawn, the water potential gradient approached the theoretical hydrostatic gradient (–0.0105 vs –0.010 MPa m–1). The gradient at solar noon was steeper (–0.0331 MPa m–1). Instantaneous stomatal conductances were not greatly different between young, sapling-sized and old-growth trees [0.094±0.033 (SD) vs 0.086±0.045 cm s–1, respectively]. Stomata of both size classes of trees appeared very sensitive to increasing vapor pressure deficits. A comparison of stable carbon isotope values from the old-growth and sapling-sized trees indicated lower stomatal conductances in the old-growth. This study provides sound documentation regarding the utility of the cohesion theory in the interpretation of water potential gradients. This study also emphasizes inherent differences between sapling-sized and tall, old-growth trees. Received: 10 January 1998 / Accepted: 12 October 1998  相似文献   

20.
In order to study protein degradation during flight in homing, a high-performance liquid chromatography technique was developed for the quantitative analysis of Nτ-methylhistidine. Secondly, it was necessary to confirm that the excretion of Nτ-methylhistidine correlates with myofilament breakdown in homing pigeons. In these experiments, ten birds were subcutaneously injected with Nτ-[14C]methylhistidine and the excreta were quantitatively collected for 1 week. Of the 94.5% radioactivity recovered, 87.1% was associated with Nτ-[14C]methylhistidine and 6.1% with N-acetyl-Nτ-[14C]methylhistidine. This rapid excretion of unmetabolized Nτ-[14C]methylhistidine validates the assumption that the amount of Nτ-methylhistidine excreted is a measure of myofilament catabolism in homing pigeons. The influence of endurance flight on protein breakdown was determined after flights from release sites 368–646 km away. Immediately after return, plasma urea and uric acid levels were increased, whereas plasma concentration of Nτ-methylhistidine remained unchanged compared to unflown control birds. Flown pigeons excreted significantly more urea and Nτ-methylhistidine within 24 h and significantly more urea and uric acid within 96 h after flight than unflown controls. Our findings support the hypothesis that in homing pigeons protein catabolism is increased during endurance flight. Elevated Nτ-methylhistidine excretion probably results from repair processes in damaged muscle fibers, including breakdown of myofilaments. Accepted: 29 October 1999  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号